Polyether-Based Lipids Synthesized with an ... - ACS Publications

The present chapter will give a short overview of the recent development ... Next to a short overview of polymeric amphiphiles currently .... Hana Bio...
0 downloads 0 Views 1MB Size
Chapter 2

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

Polyether-Based Lipids Synthesized with an Epoxide Construction Kit: Multivalent Architectures for Functional Liposomes Sophie S. Müller,1,2 Carsten Dingels,1 Anna Maria Hofmann,1 and Holger Frey*,1 1Institute

of Organic Chemistry, Johannes Gutenberg University Mainz, Duesbergweg 10-14, 55128 Mainz, Germany 2Graduate School MAINZ, Staudingerweg 9, 55128 Mainz, Germany *E-mail: [email protected].

Liposomes, vesicles consisting of phospholipids, are well known drug carriers, especially in anti-cancer treatment. Due to their improved pharmacokinetics, “stealth” liposomes, which are polymer coated vesicles, are being used in clinical applications with good results. One of the drawbacks of poly(ethylene glycol) (PEG) that is preferentially incorporated, is its lack of functional groups and its non-biodegradability. In this article new polyether-based lipids are presented that can be synthesized from an epoxide monomer library, resulting in tailored multivalent architectures. The cholesterol-based lipid-like structures offer further possibilities for functionalization, which is important for active targeting. Furthermore, a rather simple synthetic route has been developed, which leads to acid-cleavable cholesteryl PEG, thus leading to possible controlled destabilization of the liposome formulation. This process is crucial for drug release in vivo.

Introduction The formation of liposomes by self-assembly of phospholipids in water was discovered almost 40 years ago. Liposomes are colloidal vesicles consisting of a lipid bilayer with an aqueous medium interior (1). Fundamental research on the preparation, stability, release profile, encapsulation efficiency and targeting © 2013 American Chemical Society In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

of liposomes has led to clinical application. Especially for cancer therapies, the use of extremely toxic and aggressive drugs shows serious side effects. Effective delivery-systems, like liposomes, enabled an important step towards minimizing these undesired properties on healthy tissue. Advantages of these systems include a high local concentration of the anticancer drug, while protecting the body from a cytostatic drug before its release at the target site. A well known example is doxorubicin, an anthracycline, which has severe cardiotoxic effects in humans when applied directly. However, when encapsulated in lipid formulations it shows considerably increased circulation time compared to the free drug, and more importantly, lower concentration of the free drug and consequently lower cardiotoxicity. The respective product, known as Doxil™, was one of the first liposome formulations approved in the US. Conventional liposomes (see Figure 1) as lipoidal carriers have been extensively studied as drug-delivery systems, motivated by the combination of reduced side effects on healthy tissue and passive targeting (1, 2). However, the main disadvantages of such systems are the rapid removal from the blood by macrophages (mononuclear phagocyte system, MPS) after opsonin binding and uptake into the liver and spleen (3). Additionally, their physical and chemical instability results in uncontrollable properties in vivo (4, 5). To overcome these drawbacks, so-called “stealth liposomes” with surfaces modified by mainly poly(ethylene glycol) (PEG), but also polysaccharides, were developed. The presence of PEG for example effects prolonged blood circulation time (6, 7), reduced MPS uptake (8), reduced aggregation of PEGylated carriers and better (storage) stability of liposomal carriers (9).

Figure 1. left: Schematic picture of conventional liposomes and sterically stabilized liposomes, right: TEM image of liposomes containing a lipid and cholesterol-PEG.

The present chapter will give a short overview of the recent development in multifunctional “stealth” liposome preparations. This interdisciplinary area between synthetic polymer chemistry, advances in their characterization combined with new liposome preparation techniques leads to interesting new aspects in the area of “stealth” liposomes. Next to a short overview of polymeric amphiphiles currently used in “stealth” liposome preparations, the main focus will be on 12 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

polyethers such as poly(ethylene glycol) (PEG) and alternative branched and hyperbranched structures recently developed in our group. Additionally, we will give a short summary of acid-cleavable polymers in the application of degradable liposomes.

Figure 2. A selection of lipid-polymer conjugates for the incorporation into vesicle membranes to generate long-circulating “stealth” liposomes (DSPE= 1,2-Distearoyl-sn-glycero-3-phosphoethanolamine; Ch=cholesterol; PEG= poly(ethylene glycol); mPEG=methoxy poly(ethylene glycol)).

Liposome Stabilization: A Polymer Shell Is Necessary In vivo studies of conventional liposomes have shown that they are rapidly opsonized by serum proteins, and therefore taken up by cells of the mononuclear phagocyte system (MPS), such as Kupffer-cells or macrophages from the liver. This process is a key to control drug delivery, since the particles are not capable of performing their desired therapeutic task (10). Although the exact mechanism is not clearly understood yet, it is known that the type and number of proteins that attach to the vesicle’s surface can vary dramatically. Factors that influence the opsonization process include liposome size, composition and charges. Harashima et al. demonstrated a correlation between decreasing liposome size and decreasing opsonization. Furthermore, phagocytic cells remove liposomes in proportion to the amount of opsonization (11). Negatively charged liposomes increase intracellular uptake into phagocytic cells and therefore accelerate their own clearance after administration (12, 13). Since the binding of proteins 13 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

depends on a variety of physicochemical characteristics, the initial approaches to increase circulation time relied on changing these parameters. The simplest way to achieve this goal is to reduce the liposome size by sonication, extrusion or microfluidization (1). A rather recent method for the preparation of small liposomes was presented by Massing et al. They used a so-called dual asymmetric centrifuge (DAC) to produce a viscous vesicular phophoplipid gel (VPG), which can be diluted to a conventional liposome dispersion. The procedure is based on shear forces for efficient homogenization due to two rotating movements in this special centrifuge (14). The authors also showed high entrapping efficiencies of siRNA under sterile conditions (15). Another route to improve circulation time is to graft a polymer shell onto the surface of the particles (see Figure 1). The attachment of e.g., poly(ethylene glycol) (PEG), can lead to a protective, hydrophilic polymer layer, which prevents opsonin adsorption via steric repulsion. Hence, opsonization is reduced, and the probability for the uptake by the MPS system is decreased. Research groups who studied the half-life times in vivo could show extended retention periods from 5 h up to 5 days (16–19). Furthermore, it was proven that PEGylated liposomes showed improved biodistribution, which resulted in low amounts of “stealth” liposomes (10-15%) being taken up by the liver (20). Covalent attachment to lipids is clearly more stable than mere physical adsorption. Hence, “stealth” liposomes are prepared by polymer-modified lipids, which can be phospholipids or cholesterol, both natural membrane components. These polymer conjugates function as an anchor in the vesicle membrane. Selected biocompatible amphiphilic compounds used for this purpose are shown in Figure 2. Several lipid-polymer conjugates consist of phospholipids, i.e. phosphatidyl ethanolamine (PE), which is coupled to methoxy poly(ethylene glycol) (mPEG) via carbamate or amide bond formation. Furthermore, cholesterol can be used as the hydrophobic anchor. Monomethoxy poly(ethylene glycol) can be coupled to cholesteryl chloroformate by a carbonate bond (21). Our group also demonstrated an alternative method, which relies on aliphatic initiators, such as cholesterol or bis-n-hexadecyl glyceryl ether, for the ring-opening polymerization of ethylene oxide (EO). These syntheses are advantageous, since they do not require multiple reaction steps, coupling chemistry or laborious purifications steps. Furthermore, the synthesized lipids contain a hydroxyl end group, which can be used for further functionalization (22, 23).

Clinical Applications Extensive studies on the biodistribution of liposomal formulations and improved pharmacokinetic behavior of “stealth” liposomes have been done in numerous clinical tests. The best known and commercially available liposomal anti-cancer treatment, doxorubicin, is sold under the name Doxil™. The following table (Table 1) shows a selection of conventional and PEGylated liposomes for clinical usage. 14 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

15

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

Table 1. Conventional and PEGylated liposomes in clinical usage (9, 21, 24) Conventional Liposomes

PEGylated Liposomes

Drug

Company

Product name

Indication

Doxorubicin

Elan

Myocet/ Evacet

Breast cancer

Amphoterecin B

Astellas Pharma

Ambisome

Fungal infection

Daunomycin

Gilead

DaunoXome

Kaposi’s sarcoma

Vincristine

Hana Biosciences

Marqibo

Non-Hodgkin’s lymphoma

Doxorubicin

Schering Plough

Doxil/ Caelyx

Kaposi’s sarcoma, ovarian cancer

Cisplatin

Regulon

Lipoplatin

Various cancer types

Mitoxantone

Wyeth Lederle

Novantrone

Multiple sclerosis, prostate cancer, acute myeloid leukemia

In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

Liposome Stabilization: Alternatives to PEG PEG attached to lipids in pharmaceutical formulations has been investigated intensely, and these system have shown the best performance in clinical studies to date. Nevertheless, alternatives are being studied, since there is the hope for improved performance in targeting, physicochemical properties, and biocompatibility. In a recent review by Schubert et al. the disadvantages of PEG, such as its non-biodegradability, the possible degradation under stress and potential toxic side-product, were highlighted. Furthermore, the authors mention the hypersensitivity found in some cases, indicated by an activation of the complement system by PEGylated liposomes. More studies on the mechanism and the influence of other factors are necessary. Additionally, it has to be elucidated, whether other factors or the combinations of several components lead to side effects (25). General requirements for an alternative to PEG are high water-solubility, i.e., hydrophilicity of the polymer, high biocompatibility, and flexibility of the respective chain (9). Among non-biodegradable polymers poly(vinyl pyrrolidone) (PVP, commercially available) and poly(acryl amide) (PAA) have shown prolonged blood circulation times in vivo of coated liposomes (26). Poly(2-oxazoline)s are studied intensely at present, since for this type of polymer similar behavior compared to PEG has been proven (27). Figure 3 shows a selection of molecular structures of polymers currently considered for “stealth” liposome preparation.

Figure 3. Molecular structures of polymers for “stealth” liposome preparation including poly(vinyl pyrrolidone), poly(acryl amide), poly(methyl oxazoline), and poly(ethylene glycol).

In general polyethers are interesting polymers in biomedical application. Not only linear poly(ethylene glycol), but also hyperbranched structures are promising with regard to their shielding behavior in drug delivery or liposome formulations. In fact, such systems have been used as hydrophilic shells, micelles or hydrogels (28). Interestingly, surfaces covered with hyperbranched polyglycerol (hbPG) having a molecular weight around 1500-5000 g/mol showed slightly better protein repulsion than linear PEG. Presumably, the branched structure makes the polymer even more bulky, and more hydrophilic due to the high amount of hydroxyl groups, leading to a brush-like structure (29, 30). 16 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

To date, there have been very few publications on “stealth” liposomes functionalized with polyglycerols. Maruyama et al. published the synthesis and investigation of dipalmitoylphosphatidyl polyglycerols (DPP-PG), which consisted of oligomeric linear polyglycerol (lPG) attached to the phospholipid via phospholipase D (31). Effective shielding for diglycerol or tetraglycerol was observed upon using 8 mol% incorporation of the polymer-conjugate, while octaglycerol only required 4 mol% for good performance with respect to circulation times. The best result was found for DPP-hexaglycerol, which prolonged the blood circulation time, when 6 mol% were incorporated. Interestingly, such PG oligomers displayed improved performance in relation to linear PEG. Usually, PEG chains with molecular weights between 1000-5000 g/mol are used for liposomal formulations, and around 5-8 mol% is necessary to generate a considerable “stealth” effect.

Polyether-Based Lipids: Multivalent Architectures Poly(ethylene glycol) (PEG) is the most widely studied polymer for the preparation of “stealth” liposomes, which is due to its outstanding properties as a shielding layer around the vesicular carriers. Its good biocompatibility, very low toxicity as well as immunogenicity, low cost and facile coupling chemistry to hydrophobic molecules render it attractive for biomedical applications. Furthermore, its flexibility and water solubility are crucial for in vivo applications (32). However, as mentioned above, similar polyethers with different architectures exhibit the same or even better performance compared to PEG. In the following section we will highlight the recent work on polyether-based amphiphiles synthesized and characterized in our group. Phospholipids are sensitive molecules that are not stable under the very basic conditions required for the polymerization of epoxides. In general, this polymerization is an anionic ring-opening polymerization (ROP) of ethylene oxide (EO). Since the labile phospholipids are excluded under these conditions, we looked for other biocompatible options, which led us to bisalkyl glyceryl ethers and cholesterol as suitable initiators for the polymerization. These two aliphatic molecules withstand the basic ring-opening conditions as well as acidic conditions used for the deprotection of other epoxide derivatives (see Figure 4). Using a combination of ethylene oxide (EO), ethoxyethyl glycidyl ether (EEGE), isopropylidene glycidyl glyceryl ether (IGG), and glycidol as an epoxide-based construction kit, a vast variety of linear and branched architectures becomes available. Among them are complex structures, such as linear-hyperbranched amphiphiles, which combine the advantageous properties of PEG and the polyfunctionality of polyglycerol (22, 33). This aspect is one of the most important improvements compared to conventional “stealth” liposomes: the additional hydroxyl groups increase water-solubility and provide the possibility for further functionalization, such as the attachment of markers (labeling), antibodies or targeting groups. Maximum biocompatibility is achieved by using cholesterol as the initiator and hbPG as a polymer with excellent biocompatibility (28, 34–36). 17 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Synthesis of Multivalent Architectures

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

The synthesis of multivalent lipids is described with cholesterol as an initiator, since the resulting polymer is expected to show very good biocompatibility. Cholesterol is a natural membrane component and also commercially available. In the lipid structures prepared, cholesterol can function as the membrane anchor.

Figure 4. Reaction sequence for the synthesis of cholesterol initiated polyethers: Ch-PEG, Ch-lPG, Ch-lPGG.

The first step is the formation of the cholesterol alkoxides, which represent the initiator for the subsequent polymerization. The degree of deprotonation employed is 90% in the case of CsOH as the deprotonating agents. Due to the rapid proton exchange between cholesterol and the growing chain, almost 100% of the initiator molecules are incorporated into the resulting polyether amphiphile. To form the linear polymer, ethylene oxide can be polymerized using the standard oxyanionic ring-opening polymerization technique (37). In order to obtain linear polyglycerol protected epoxide derivatives such as ethoxyethyl glycidyl ether (EEGE) or isopropylidene glycidyl glyceryl ether (IGG) can be used. Deprotection of the acetal groups leads to one (EEGE) or two (IGG) hydroxyl groups per monomer unit. Using this protocol, multifunctional, linear polyglycerol can be synthesized in a one-pot approach (Figure 4). Using a two-step procedure, it is possible to tailor hyperbranched structures based on linear macroinitiators. Glycidol is polymerized by the ring-opening multibranching technique in a slow monomer addition step (Figure 5). The multihydroxy-precursor polymer is crucial for the “hypergrafting” of glycidol since excellent conditions and low polydispersities are required. Usually the degree of deprotonation is around 25%, which allows good control over the anion concentration and prevents homopolymerization of glycidol. For the slow monomer addition step a syringe pump is used, which adds the monomer in low concentration over a certain amount of time (around 18-24 h, depending on the batch size). 18 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

Figure 5. Reaction sequence for the synthesis of cholesterol initiated hyperbranched structures (Ch-hbPG).

Following the above mentioned reaction sequence, one can combine any epoxide derivative in the way of a molecular construction kit, permitting rapid access to polyfunctional lipids via ROP (see Figure 6). Random copolymerization allows for the synthesis of Ch-PEG-co-PGG or Ch-PEG-co-lPG using EO and IGG or EEGE, respectively. Additionally, linear hyperbranched structures are available via the ROP of ethylene oxide followed by EEGE or IGG, the deprotection step and subsequent slow monomer addition of glycidol onto the macroinitiator polymer. The resulting architectures show low to moderate polydispersities (PDI= 1.1-1.6) and can be characterized by size exclusion chromatography (SEC), NMR spectroscopy or matrix assisted laser desorption time-of-flight mass spectrometry (MALDI ToF MS). Furthermore the critical micelle concentration was determined to be in the range of 1.4-40.7 mg/L, depending on the molecular structure (33).

Figure 6. Possible lipid architectures available by using EO, EEGE, IGG, and glycidol for the ring-opening polymerization initiated by either cholesterol or bis-n-hexadecyl glyceryl ether. 19 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

In a recent work by Kressler et al. the linear poly(ethylene glycol)30-b-hbPG24 copolymer was investigated in mixed layers with 1,2-dipalmitoyl-sn-glycero-3phosphocholine (DPPC). In adsorption measurements it was demonstrated that an intense affinity of the amphiphilic block copolymer to DPPC is given after the injection of the polymer into the water subphase. The surface pressure was determined to be 48.2 mN/m, showing the fast penetration of the hydrophobic cholesterol into the lipid monolayer, as well as good interaction with DPPC. Figure 7 shows the DPPC monolayer at the air-water interface with the amphiphilic polymer, which is attached to the lipid layer via its cholesterol moiety (38).

Figure 7. DPPC monolayer and adsorbed cholesterol moiety of Ch-PEG-hbPG to the air-water interface. (Reproduced with permission from reference (38). Copyright 2012 Springer Verlag.)

Several multifunctional polymers have been synthesized, showing generally good adsorption behavior at DPPC monolayers. In future work, the properties in liposomal formulations in vivo will be studied. These novel coatings combine the advantages of PEG as well as PG and might increase liposome stability. Furthermore, it is important to target the vesicles, which can be easily accomplished by the functionalization of peripheral hydroxyl groups. Model reactions, such as the attachment of the dye rhodamine B by click chemistry demonstrated the utility of the branched structure (33). The performance in blood serum is currently under investigation, and further experiments in vivo will be carried out in the near future.

Liposome Destabilization: pH-Sensitive Lipids with an Acid-Labile Moiety Acid-sensitive PEG amphiphiles, such as lipids, have attracted considerable attention in biomedical applications, since controlled destabilization of liposomes is necessary for the release of the drug incorporated in the vesicular transporter. 20 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

Those “stealth” systems are usually stable at neutral pH, which translates to stability during the circulation in the body, whereas destabilization is needed for the fusion with membranes or the escape from endosomal vesicles (39). The fusion and the release of the drug in slightly acidic tissue, such as tumor tissue or inflammatory tissues, can be realized with pH-sensitive PEG coated liposomes, where the protecting shell can be shed at a pH around 6.5 for tumor tissues. One major advantage is that no external stimulus is required for triggering the drug release. In this section we would like to focus on acid-sensitive PEG lipid analogs that have been used for liposome preparation. Furthermore, we present very recent work that has been carried out in our group, which is based on an acid-labile cholesterol derivative as the initiator for the ring-opening polymerization of ethylene glycol. Different approaches have been developed to tune the pH-sensitivity of the polymer shell around nanoparticles or liposomes. Recently, Clawson et al. have published the synthesis and characterization of lipid-polymer hybrid nanoparticles with a PEG shell that sheds under acidic conditions. They used a lipid-(succinate)mPEG as the amphiphilic polymer, which was synthesized via the coupling of 1,2-dipalmitoyl-sn-glycer-3-phospho(ethylene glycol) and methoxy poly(ethylene glycol), endfunctionalized with succinate prior to the coupling step (40). In 2003 an acid-labile poly(ethylene glycol) PEG-conjugated lipid, (R)-1,2-di-O-(1‘Z,9‘Z-octadecadienyl)-glyceryl-3-(ω-methoxy-poly(ethylene glycolate, MW5000) (BVEP), was synthesized and mixed with 1,2-dioleoylsn-glycero-3-phosphoethanolamine (DOPE) in liposomes to investigate destabilization and membrane-membrane fusion after acid-catalyzed hydrolysis of the vinyl ether linkages. The research group showed that PEG removal occurred after hydrolysis, but that it resulted in undesired payload leakage and liposome collapse as well (41). Using cholesterol as the lipophilic part, Boomer et al. presented a six-step synthesis for the preparation of cholesterol-vinyl ether-PEG conjugates (CVEP), which degraded under mildly lowered pH values. Cleavage resulted in PEG removal, leading to content release and thus an increased bioavailability of a potent drug (42). In our group a cholesterol derivative was used directly as the initiator for the ring-opening polymerization of ethylene oxide (EO) (43). The addition of the steroid (1) to 2-acetoxyethyl vinyl ether (AcVE, 2) was carried out in dichloromethane, catalyzed by p-toluene sulfonic acid, leading to acetoxyethyl 1-(cholesteryloxy)ethyl ether (3). The advantages of AcVE are the following: The acetate group can be used as a protection group and is removed under basic conditions with little effort, leading to glycol 1-(cholesteryloxy)ethyl ether (4). Hence, the hydroxyl group, which is important for the oxyanionic polymerization, is liberated easily. This group is used as an alkoxide after deprotonation and is structurally related to the growing chain end, which results in good initiation kinetics. Hence, it was possible to synthesize a scissile initiator for the polymerization of EO in three steps (Figure 8). The polymer (5) was characterized by NMR spectroscopy, SEC, and MALDI-ToF MS. 21 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

Figure 8. Synthesis of the pH-sensitive cholesterol derivative for the initiation of ethylene oxide. The insertion of such a scissile hydrophobic unit leads to responsive materials that are cleaved under acidic condition leading to a loss of their amphiphilicity. This is a desired feature, when it comes to destabilizing the “stealth” liposomes. Acid-sensitivity was demonstrated as a proof of principle in a UV-Vis spectrometer. An aqueous solution of the scissile lipid analog was acidified by adding hydrochloric acid. After a while the solutions turned turbid, as the released cholesterol precipitated and the transmission began to decrease. Almost all light was scattered in the end, due to cholesterol precipitation. As expected, cholesterol was released faster at higher reaction temperature (T=25 °C vs. T=37 °C), as indicated by the shorter initial phase and more negative slope of the trace. To confirm complete removal of the steroid, a similar experiment was performed with the scissile cholesteryl PEG, in which the precipitated cholesterol and the aqueous solution were separated and 1H spectra were recorded. A clean cholesterol spectrum was obtained, whereas the aqueous phase exhibited pure PEG diol. Hence, the cholesteryl initiator was released completely under these reaction conditions. This system represents a novel approach towards cleavable amphiphiles, which are promising for “stealth” liposome preparation. One acid-sensitive moiety can be cleaved resulting in biocompatible cholesterol and PEG, respectively. The molecular weight of the PEG polymer chain is typically around 2000-5000 g/mol, so the polymer is small enough to be excreted by the kidney and be eliminated from the body. We believe this approach to be promising in shedding the protecting layer, and thus destabilizing liposomes at the site of action, where lower pH-values are present. Further studies are planned to synthesize acid-cleavable, hyperbranched lipids for multifunctional liposomes.

Conclusion Many approaches have been made in the development of effective drug delivery systems, especially in the field of liposomes. The vesicles used as a transporter for hydrophilic or hydrophobic drugs suffered from several disadvantages such as low stability and low circulation times in vivo. One of the biggest steps towards long-circulating liposomes was the incorporation of 22 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

a biocompatible, water-soluble polymer, such as poly(ethylene glycol) (PEG). The improved performance of the liposomal formulation resulted from the stabilizing polymer shell around the vesicle, which protects the drug carrier and additionally prevents opsonin adsorption. These unique properties lead to reduced blood clearance and hence prolonged blood circulation times. Additionally, the enhanced accumulation and retention effect (EPR) helps to increase drug concentration at the site of action. It is no surprise that liposomal formulations, with or without a polymer shell, are used in clinical applications, especially in cancer treatment, due to their outstanding pharmacokinetics. PEG coated liposomes, so-called “stealth liposomes” are applied in various treatments such as Kaposi’s sarcoma, breast cancer or fungal infections. Nevertheless, new synthetic polymers offer various advantages that may further enhance the applicability of “stealth” liposomes. Among these are biodegradability and multifunctionality, which is important for the attachment of markers and targeting moieties. In this context we presented new polymeric amphiphiles based on poly(ethylene glycol) and polyglycerol (linear and hyperbranched) that represent a new class of multifunctional lipid analogs with various architectures. The multiple hydroxyl groups increase aqueous solubility of the polymer and offer the possibility for effective targeting through further functionalization. In addition, the degradability of acid-sensitive Ch-PEG polymers was discussed, in contrast to chemically inert PEG, resulting in possible destabilization of the “stealth” liposome in the acidic tumor tissue or acidic endosomes. The desired destabilization is still one of the key tasks that remain problematic in liposome research. Regarding the advantages of the hyperbranched lipids, we believe that this class of lipids is promising with respect to drug delivery and “stealth” components. Investigations on the stealth effect in vivo, as well as monolayer studies, are currently in progress.

References 1. 2. 3. 4. 5. 6.

7.

8. 9.

Sharma, A.; Sharma, U. S. Int. J. Pharm. 1997, 154 (2), 123–140. Lian, T.; Ho, R. J. Y. J. Pharm. Sci. 2001, 90 (6), 667–680. Lasic, D.; Papahadjopoulos, D. Science 1995, 267 (5202), 1275–1276. Allen, T. M.; Chonn, A. FEBS Lett. 1987, 223 (1), 42–46. Woodle, M. C.; Lasic, D. D. Biochim. Biophys. Acta 1992, 1113 (2), 171–199. Woodle, M. C.; Matthay, K. K.; Newman, M. S.; Hidayat, J. E.; Collins, L. R.; Redemann, C.; Martin, F. J.; Papahadjopoulos, D. Biochim. Biophys. Acta 1992, 1105 (2), 193–200. Torchilin, V. P.; Omelyanenko, V. G.; Papisov, M. I.; Bogdanov, A. A.; Trubetskoy, V. S.; Herron, J. N.; Gentry, C. A. Biochim. Biophys. Acta, Biomembr. 1994, 1195 (1), 11–20. Blume, G.; Cevc, G. Biochim. Biophys. Acta, Biomembr. 1990, 1029 (1), 91–97. Immordino, M. L.; Dosio, F.; Cattel, L. Int. J. Nanomed. 2006, 1 (3), 297–315. 23

In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

10. Owens, D. E.; Peppas, N. A. Int. J. Pharm. 2006, 307 (1), 93–102. 11. Harashima, H.; Sakata, K.; Funato, K.; Kiwada, H. Pharm. Res. 1994, 11 (3), 402–406. 12. Lee, K. D.; Hong, K.; Papahadjopoulos, D. Biochim. Biophys. Acta 1992, 1103 (2), 185–197. 13. Gabizon, A.; Papahadjopoulos, D. Biochim. Biophys. Acta 1992, 1103 (1), 94–100. 14. Massing, U.; Cicko, S.; Ziroli, V. J. Controlled Release 2008, 125 (1), 16–24. 15. Hirsch, M.; Ziroli, V.; Helm, M.; Massing, U. J. Controlled Release 2009, 135 (1), 80–88. 16. Klibanov, A. L.; Maruyama, K.; Torchilin, V. P.; Huang, L. FEBS Lett. 1990, 268 (1), 235–237. 17. Senior, J.; Delgado, C.; Fisher, D.; Tilcock, C.; Gregoriadis, G. Biochim. Biophys. Acta, Biomembr. 1991, 1062 (1), 77–82. 18. Papahadjopoulos, D.; Allen, T. M.; Gabizon, A.; Mayhew, E.; Matthay, K.; Huang, S. K.; Lee, K. D.; Woodle, M. C.; Lasic, D. D.; Redemann, C.; Martin, F. J. Proc. Natl. Acad. Sci. U.S.A. 1991, 88 (24), 11460–11464. 19. Stathopoulos, G. P.; Boulikas, T.; Vougiouka, M.; Deliconstantinos, G.; Rigatos, S.; Darli, E.; Viliotou, V.; Stathopoulos, J. G. Oncol. Rep. 2005, 13 (4), 589–595. 20. Allen, T. M.; Hansen, C.; Martin, F.; Redemann, C.; Yau-Young, A. Biochim. Biophys. Acta, Biomembr. 1991, 1066 (1), 29–36. 21. Zhao, X. B.; Muthusamy, N.; Byrd, J. C.; Lee, R. J. J. Pharm. Sci. 2007, 96 (9), 2424–2435. 22. Hofmann, A. M.; Wurm, F.; Hühn, E.; Nawroth, T.; Langguth, P.; Frey, H. Biomacromolecules 2010, 11 (3), 568–574. 23. Reuter, S.; Hofmann, A. M.; Busse, K.; Frey, H.; Kressler, J. Langmuir 2010, 27 (5), 1978–1989. 24. Torchilin, V. P. Nat. Rev. Drug Discovery 2005, 4 (2), 145–160. 25. Knop, K.; Hoogenboom, R.; Fischer, D.; Schubert, U. S. Angew. Chem., Int. Ed. 2010, 49 (36), 6288–6308. 26. Torchilin, V. P.; Shtilman, M. I.; Trubetskoy, V. S.; Whiteman, K.; Milstein, A. M. Biochim. Biophys. Acta, Biomembr. 1994, 1195 (1), 181–184. 27. Zalipsky, S.; Hansen, C. B.; Oaks, J. M.; Allen, T. M. J. Pharm. Sci. 1996, 85 (2), 133–137. 28. Calderon, M.; Quadir, M. A.; Sharma, S. K.; Haag, R. Adv. Mater. 2010, 22 (2), 190–218. 29. Siegers, C.; Biesalski, M.; Haag, R. Chem.—Eur. J. 2004, 10 (11), 2831–2838. 30. Kainthan, R. K.; Zou, Y.; Chiao, M.; Kizhakkedathu, J. N. Langmuir 2008, 24 (9), 4907–4916. 31. Maruyama, K.; Okuizumi, S.; Ishida, O.; Yamauchi, H.; Kikuchi, H.; Iwatsuru, M. Int. J. Pharm. 1994, 111 (1), 103–107. 32. Roberts, M. J.; Bentley, M. D.; Harris, J. M. Adv. Drug Delivery Rev. 2002, 54 (4), 459–476. 24 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by UNIV OF SYDNEY on February 2, 2014 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch002

33. Hofmann, A. M.; Wurm, F.; Frey, H. Macromolecules 2011, 44 (12), 4648–4657. 34. Wilms, D.; Wurm, F.; Nieberle, J.; Boehm, P.; Kemmer-Jonas, U.; Frey, H. Macromolecules 2009, 42 (9), 3230–3236. 35. Wilms, D.; Stiriba, S.-E.; Frey, H. Acc. Chem. Res. 2010, 43 (1), 129–141. 36. Kainthan, R. K.; Gnanamani, M.; Ganguli, M.; Ghosh, T.; Brooks, D. E.; Maiti, S.; Kizhakkedathu, J. N. Biomaterials 2006, 27 (31), 5377–5390. 37. Mangold, C.; Wurm, F.; Frey, H. Polym. Chem. 2012, 3 (7), 1714–1721. 38. Peng, X.; Hofmann, A.; Reuter, S.; Frey, H.; Kressler, J. Colloid. Polym. Sci. 2012, 290 (7), 579–588. 39. Hong, R. L.; Huang, C. J.; Tseng, Y. L.; Pang, V. F.; Chen, S. T.; Liu, J. J.; Chang, F. H. Clin. Cancer Res. 1999, 5 (11), 3645–3652. 40. Clawson, C.; Ton, L.; Aryal, S.; Fu, V.; Esener, S.; Zhang, L. Langmuir 2011, 27 (17), 10556–10561. 41. Boomer, J. A.; Inerowicz, H. D.; Zhang, Z. Y.; Bergstrand, N.; Edwards, K.; Kim, J. M.; Thompson, D. H. Langmuir 2003, 19 (16), 6408–6415. 42. Boomer, J. A.; Qualls, M. M.; Inerowicz, H. D.; Haynes, R. H.; Patri, V. S.; Kim, J.-M.; Thompson, D. H. Bioconjugate Chem. 2009, 20 (1), 47–59. 43. Dingels, C.; Müller, S. S.; Steinbach, T.; Tonhauser, C.; Frey, H. Biomacromolecules 2013 DOI:10.1021/bm3016797.

25 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.